* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download Reversible lithium uptake by CoP3 at low potential: role of the anion
Survey
Document related concepts
Transcript
Electrochemistry Communications 4 (2002) 516–520 www.elsevier.com/locate/elecom Reversible lithium uptake by CoP3 at low potential: role of the anion V. Pralong, D.C.S. Souza, K.T. Leung, L.F. Nazar * Department of Chemistry, University of Waterloo, 200 University Ave. W., Waterloo, Ont., Canada N2L 3G1 Received 25 April 2002; accepted 3 May 2002 Abstract Although various transition metal compounds in Group V (nitrides and antimonides) have been reported to act as low potential Li insertion hosts, the phosphides have remained unexplored to date. We show here that lithium uptake and extraction in the metal phosphide, CoP3 , provides a reversible capacity of 400 mAh/g at an average potential of 0.9 V vs Li=Liþ , via a novel mechanism. The latter was revealed using a combination of X-ray diffraction, electron microscopy, and potentiodynamic and galvanostatic intermittent measurements coupled with X-ray photoelectron spectroscopy (XPS). Initial uptake of Li forms highly dispersed cobalt clusters embedded in a matrix of Li3 P; extraction of Li from this ion-conductive matrix on charge yields nano-particles of LiP, with little change evident in the oxidation state of the Co site. This shows that contrary to the case of metal oxides here the anion plays the major role in reduction and oxidation. We expect this is a general phenomenon for phosphides, and to be of fundamental interest and future importance in the search for new negative electrode materials. Ó 2002 Published by Elsevier Science B.V. Keywords: Cobalt phosphide; Anode; Lithium-ion battery; Transition metal pnictides; XPS; TEM 1. Introduction The development and study of new negative electrode materials for rechargeable lithium batteries remains a major topic in the area of the energy storage. Alternatives to commercial carbonaceous materials are eagerly sought to overcome the limitations of low gravimetric and volumetric capacity and safety concerns that these materials exhibit, and to match the higher capacities of new emerging cathode materials. Meeting the requirements of reversibility and low cost is difficult, however. Explorations of transition metal oxides as anode materials, first reported by Idota et al. [1], Nazar and coworkers [2] and by Tarascon and coworkers [3], developed new concepts concerning the mechanism of Li ‘‘insertion’’. For example, in bulk MoO3 , Li uptake drives the reduction of the metal to form a nano-phase matrix of Li2 O, and metal suboxide particles: on charge, Mo-oxide is regenerated, although ‘‘crystalline MoO3 ’’ is not fully achieved. Similarly, reduction of nano-phase * Corresponding author. Tel.: +1-519-888-4637; fax: +1-519-7460435. E-mail address: [email protected] (L.F. Nazar). 1388-2481/02/$ - see front matter Ó 2002 Published by Elsevier Science B.V. PII: S 1 3 8 8 - 2 4 8 1 ( 0 2 ) 0 0 3 6 3 - 6 CoO produces Co nano-particles in a matrix of Li2 O; and nano-CoO is regenerated on charge [4]. It has subsequently been revealed that the latter reaction can be thought of as a displacement reaction of the Co2þ ions within the oxide matrix for Liþ on discharge, with the reverse occurring on charge [5]. In these relatively ‘‘ionic’’ systems, reduction is driven by the difference in thermodynamic free energy between the oxidized and reduced states of the metal. Recent interest has also been piqued by reports of Li uptake in transition metal compounds of Group V, such as nitrides [6] and antimonides [7], which show a lower intercalation potential compared to the respective oxides. This results from the lower formal oxidation state of the metal, and strong covalent character of the Mpnictogen bond, leading to high lying mixed anion-metal bands, and a high degree of electron delocalization [8]. Where a fully lithiated end member exists, conversion between two crystalline phases can occur on lithium insertion, i.e., MnP4 $ Li7 MnP4 [9]. Here, we report a completely different reversible Li uptake in metal phosphides exemplified by CoP3 . We have found unique behavior in this system: namely the redox behavior is based on shuttling between two different phosphorus oxidation states. V. Pralong et al. / Electrochemistry Communications 4 (2002) 516–520 517 2. Experimental 3. Results and discussion Synthesis of the cobalt phosphide was performed by heating stoichiometric amounts of metallic cobalt and red phosphorus at 650 °C for 24 h, in a stainless steel tube sealed under argon. Powder diffraction patterns were obtained using a Siemens D500 diffractometer equipped with a diffracted beam monochromator (CuKa radiation). The starting material, CoP3 was a pure skutterudite phase, as confirmed by Rietveld refinement of its XRD pattern that yielded values (cubic space ) as previously reported [10]. group IM3; a ¼ 7:7082 A The structure consists of CoP6 octahedra corner-shared to form P4 rings (Fig. 1, inset). The Li uptake properties of the material were evaluated in Swagelok-type cells vs the Li=Liþ couple using lithium as the anode. Mixtures of the active material and 40 wt% carbon black were cast on nickel current collectors. The cells were assembled in an argon filled dry-box using a 1 M solution of LiPF6 in 1:1 EC/DMC as the electrolyte, and examined using a MacPile controller (Biologic S.A., Claix, France). X-ray photoelectron spectrocopy (XPS) analysis was performed using a VG Scientific XPS Microprobe ESCALab 250 (with focused monochromatized AlKa radiation of energy 1486.6 eV, with a chamber pressure <1010 mbar) on an irradiated area of 0:4 1 mm2 . The cycled samples were opened in the dry-box, washed with acetonitrile and transferred to the XPS analysis chamber through a sample load-lock system to prevent air exposure. Charge compensation was employed during data collection using a combination of a low energy flood gun and low energy ions (care was taken to avoid charging). These samples were also analyzed by means of transmission electron microscopy (TEM, Philips CM12). The voltage vs composition curves for the insertion of Li into CoP3 , acquired at a rate of 1Liþ per 10 h are shown in Fig. 1. The first discharge process is characterized by a flat plateau corresponding to the uptake of 9 Li (after correction for the carbon contribution; the initial feature in the curve at 0.8 V corresponds to the SEI formation with carbon). On charge, 6 Li are extracted, leading to a reversible capacity of 1000 mAh/g of CoP3 , that fades to 600 mAh/g after 10 cycles and stabilizes at > 400 mAh=g. By means of potentiodynamic cycling with galvanic acceleration, a bell shaped current response associated with a constant voltage (0.35 V; not shown here) during the first discharge shows that the plateau corresponds to a simple two-phase decomposition. We see no evidence for the formation of intermediate ternary phases ‘‘Lix CoPy ; these are also not expected since only a lithium-poor/metal-rich phase, LiCo6 P4 , exists in the Li–Co–P phase diagram. The reaction of CoP3 with 9 Liþ that we propose on discharge: (CoP3 þ 9Liþ þ 9e ! 3Li3 P þ \Co") is comparable to behavior reported by Tirado and coworkers [7] for CoSb3 , which has been proposed to transform into ðLi3 Sb þ CoÞ on Li uptake. In the case of CoP3 , the XRD pattern product at the end of the first discharge (Fig. 2) clearly shows the reflections of the Li3 P phase but not those of metallic cobalt. Any Co particles must be of nano-sized dimensions (well below the limit of detection in both XRD and TEM – see later). The charge process, corresponding to the extraction of 6 Li, and subsequent cycling behavior is less obvious. The material becomes quasi-amorphous on oxidation, Fig. 1. Potential/composition curve on the first galvanostatic cycling of CoP3 –Li=Liþ cells performed at a nominal C/10 rate in the 1.7–0.2 V potential window. Inset: Projection of the CoP6 octahedra along the [0 0 1] direction in the structure of CoP3 : Fig. 2. Ex situ XRD patterns of CoP3 at different cycling stages: (a) before cycling; (b) discharge at 0.1 V after 1 cycle, Miller indices (*) mark the position of Li3 P; (c) charge at 1.7 V after 1 cycle, reflections mark the positions expected for LiP. The reflection of the cell holder due to Al are indicated; ¼ unidentified phase arising from reaction with the electrolyte. 518 V. Pralong et al. / Electrochemistry Communications 4 (2002) 516–520 as shown by X-ray diffraction and only very weak, broad reflections are evident even after data collection for extended periods (Fig. 2) [11]. An important question is whether the cationic or the anionic network plays the major role in the electrochemical redox process. That is, does oxidation result in re-formation of highly dispersed CoP3 (as in the case of the oxide, CoO) or is the process more complex? To more closely investigate the lithium-driven structural and morphological changes, we examined the CoP3 electrode at various stages of reduction and oxidation processes by means of transmission electron microscopy and X-ray photoelectron spectroscopy. The TEM micrograph of the starting material (Fig. 3(a)) shows it consists of particles of dimensions ranging from 200 to 100 nm; as expected, selected area electron diffraction (SAED) patterns taken down the [1 1 1]* zone axis (Fig. 3(a), right) are consistent with the presence of well-crystallized particles in accord with the XRD results. When the CoP3 is fully reduced (Fig. 3(b)), we observed a complete change of the texture compared to the starting material. The bright field image reveals the (a) (b) (c) Fig. 3. TEM images and corresponding SAED patterns on the right, of the (a) starting material CoP3 , (b) fully discharged state, with the reflections of Li3 P indicated; (c) fully charged state showing the reflections of LiP – the inset shows the ‘‘halo’’ SAED image taken with a normal diaphragm aperture, similar to that in (b). particles embedded within the presence of 10–20 A crystallite, and preservation of the overall shape of the starting particle. In this pseudomorphic reaction, a partially amorphous material is formed, as confirmed by the SAED pattern that consists of broad, weak rings. The reflections were indexed to Li3 P (Table 1) in accord with the XRD patttern. Analysis of the broad ring pattern did not reveal any interplanar spacings assignable to cobalt crystallites (Table 1 and Fig. 3(b), right), indicating that the metal must be very highly dispersed. It is also possible that is weakly bonded to the matrix as highly metal-rich phosphide clusters. Particle dimen would not be observable by SAED due sions below 10 A to extreme broadening. On charge, the bright field image shows the pseudoamorphous character of the active material is preserved although the particle size is more than halved (Fig. 3(c)). Nonetheless, the weak, broad reflections observed in the XRD pattern are resolvable, and identifiable by SAED in the TEM owing to the smaller crystallite size that can be probed by this method. The crystallites yielded a very broadened SAED pattern consistent with their reduced dimensions (Fig. 3(c), right, inset) in which only a few concentric diffuse ‘‘halo’’ rings were visible, necessitating employment of a larger diffraction diaphragm to improve the image quality 1; 2 [12,13]. The resulting SAED pattern (Fig. 3(c), right) was best indexed to LiP (Table 1), but not to CoP3 . 3 As no lines can be ascribed to Li3 P, it has been consumed during oxidation (Table 1). Hence, the electrochemical process appears best characterized as a redox reaction between Li3 P and LiP. Evidence for the reversibility, and nature of this intriguing process was supplanted by XPS measurements that provide a direct ‘‘finger print’’ of the sample. The binding energy is characteristic of the elements present and their chemical states. Unfortunately, the resolvable Co2p XPS region cannot be used to elucidate the exact role of the cobalt, since metallic and covalent cobalt 2p ð1=2; 3=2Þ states are at almost the same energy level; furthermore the binding energy threshold does not 1 The diffraction diaphragm was increased by a factor of 10 to increase the data sampling area: this explains the brightness of the SAED image associated with a quasi-amorphous phase in the relatively featureless XRD pattern. This technique is commonly used for ‘‘amorphous’’ silicon, for example, where similar effects are seen for nano-sized particles that do not show discrete features in their XRD pattern [12]. 2 The co-existence of an ‘‘amorphous’’ XRD pattern, and an indexable ring pattern in the SAED using a standard aperture has also been observed; see [12]. 3 Reflection positions matched those of LiP: P21/c, a ¼ 5:582(1) b ¼ 4:94(1) c ¼ 10:255(2), b ¼ 118:150, JCPDF no. 83–1575. Small deviations may result from the presence of defect-laden LiP, owing to its nano-dimensions and method of formation. Note that the two strongest lines of CoP3 : (2 0 0) and (2 2 0), do not match the observed lines, Table 1. V. Pralong et al. / Electrochemistry Communications 4 (2002) 516–520 519 Table 1 List of expected and observed d-spacings from XRD and SAED patterns for the products after discharge and charge Li3 P, calculated ) reflections (A d-spacing, SAED ) pattern (A d-spacing, XRD ) pattern (A (a) discharged state 3.69 (1 0 1) 2.14 (1 1 0) 1.86 (1 1 2) 1.42 (2 1 0) 1.22 (3 0 0) 3.69 2.14 1.85 1.40 1.23 3.5 2.1 1.8 – – LiP, calculated ) d-reflections (A d-spacing, SAED ) pattern (A Approximate ) d-spacing (XRD) (A CoP3 , calculated reflections ) near observed (A (b) charged state 3.34 (0 1 2) 2.94 (1 1 1) 2.46 (2 0 0) 2.27 ()1 1 4) 3.40 2.94 2.46 2.21 3.4 3.0 – – 3.85 2.72 2.44 2.22 change significantly from the pristine material upon reduction or oxidation (Co2p1=2;3=2 at 793.5/778.4 eV in CoP3 ; 794.0/779.0 eV on discharge; at 793.7/778.7 eV on charge). Examination of the overlapping Co3p and Li1s region is more revealing, however: Fig. 4 shows the spectra of the fully charged and discharged electrodes, after ion sputtering to clean the surface. In both spectra, the broad feature (P1) at 60 eV is assigned to the Co3p ð1=2; 3=2Þ states, by comparison with reference data [14]. This peak should, of course, remain unchanged in accord with the invariance of the Co3p states (see above). The source of the change in the spectra therefore must lie in the Li 1s states. In both spectra, the signature at 56 eV (P2) corresponds to Liþ from the residual electrolyte (LiPF6 ) [14]. In the spectrum of the discharged material (Fig. 4(a)), this peak is overlapped with (a) P1 P2,3 (b) P1 P3 P2 2.05 (1 1 1) 1.78 (2 0 0) (2 0 0) (2 2 0) (0 1 3) (2 2 2) another at 56.87 eV, but is distinguishable in a curve deconvolution. This peak (P3) is assigned to the Li 1s state in Li3 P, in good accordance with the literature value of 56.9 eV [14]. It shifts to a higher binding energy (58.2 eV) on charge (Fig 4(b)), indicating the phosphorous oxidation state decreases. This peak is assigned to LiP by comparison with its XPS spectrum (binding energy: 59.03 eV). 4 A simplified mechanism accounting for the initial reversible capacity of 6 Li can therefore be written as follows: First discharge process : CoP3 þ 9Liþ þ 9e ! 3Li3 P þ \Co" Subsequent cycling : 3Li3 P () 3LiP þ 6Liþ þ 6e In this system, the relatively high electronic conductivity of CoP3 enhances initial electronic transport within the electrode material and facilitates the decomposition process. After discharge, the high ionic conductivity of Li3 P of 103 S/cm (that has spurred its investigation as a solid electrolyte [15]) promotes Li ion transfer within the composite matrix. The role of the cobalt in the redox process (if any) is still unclear; since the cobalt is in a highly dispersed state in both the discharged and charged materials which preclude investigation of its nature, we can only conclude that it acts as an ‘‘active spectator’’ after initial discharge as bulk metallic cobalt is not formed. This is also in accordance with the high levels of carbon that are needed in the compound electrode to promote electronic conductivity. The high degree of Co dispersion and bonding to the Li– 4 Fig. 4. XPS Lithium-1s spectra of Lix CoP3 in its (a) fully discharged and (b) fully charged states. Co metal, calculated reflections ) near observed (A This spectrum was collected on the reference material, LiP, synthesized from heating (Li0 þ red P) at 600 °C for 24 h in a sealed stainless steel tube. Note that the binding energy of bulk LiP and that prepared in situ differ by about 0.2 eV, consistent with the shift expected for nano-particles in the quantum size regime. 520 V. Pralong et al. / Electrochemistry Communications 4 (2002) 516–520 P matrix may nonetheless enhance Li transport by grain boundary effects. The fact that neither CoP3 nor an amorphous form of lithium cobalt phosphide are regenerated on charge sets this mechanism apart from transition metal oxides, where partial or complete recuperation of the oxide structure occurs. This can be understood since the reformation of CoP3 nano-crystallites is unlikely in the more covalently bonded phosphides, owing to the thermodynamics of the Li–Co–P system which should favor LiP formation on charge as a result of its high lattice energy. Nonetheless, the fully reduced material can be considered as a highly dispersed, phase-separated metastable analog of ‘‘Li6 CoP3 ’’. In this respect, the process is different from Li uptake in phosphides where ternary lithiated phases do exist, such as MnP4 , In this case, reversible conversion to crystalline Li7 MnP4 results [9]. Both processes, however, are initiated by electron uptake that primarily occurs at the anion centre, giving rise to scission of the P–P bonds. 4. Conclusions This work highlights the importance of the anion in the electrochemical process, proving that the redox active centers are not necessarily cationic in transition metal compounds but can be based almost uniquely on the anionic network. In this regard, it is worthwhile recalling that the anion alone also acts as the redox center in the well-known Li–sulfur battery. This cell is based on reactions of Li and sulfur that reversibly yield Li2 S. Development of those cells has been hampered by the high ionic and electronic resistivity of sulfur, solubility of polysulfide products, and other factors [16]. Investigations of transition metal pnictides as anodes for Li-ion batteries may re-open new opportunities in this fascinating area. Such studies would also focus on the aspect of microstructural control that would lead to improve cycling stability in these materials, work that is currently in progress in our lab. Acknowledgements The authors are grateful to B. Beaudoin (Amiens), M.M. Treacy (NEC labs, Princeton), and D. Perovic (U of Toronto), and for insightful discussions on TEM. D.K. Fujita and Dr. Enke are thanked for their help. D.C. Souza thanks CAPES/Brazil for their financial support in the form of a scholarship, and LFN gratefully acknowledges the support of NSERC for funding this work. References [1] Y. Idota, US Patent, 5478671 (1995); H. Tomyama, Jap. Patent, 07-029608 (1995); C. Sigala, D. Guyomard, Yves Piffard, M. Tournoux, C. R. Acad. Sci. Paris 320 (1995) 523. [2] F. Leroux, G.R. Ouvrard, W.P. Power, L.F. Nazar, Electrochem. Solid State Lett. 1 (1998) 255; F. Leroux, L.F. Nazar, New Mater. Batteries Fuel Cells II 55 (2000) 173. [3] S. Denis, E. Baudrin, F. Orsini, G. Ouvrard, M. Touboul, J.-M. Tarascon, J. Power Sources 81 (1999) 79. [4] P. Poizot, S. Laruelle, S. Grugeon, L. Dupont, J.M. Tarascon, Nature 407 (2000) 496. [5] M.N. Obrovac, R.A. Dunlap, R.J. Sanderson, J.R. Dahn, J. Electrochem. Soc. 148 (2001) A576. [6] M. Nishijima, T. Kagohashi, M. Imanishi, Y. Takeda, O. Yamamoto, S. Kondo, Solid State Ionics 83 (1996) 107; T. Shodai, S. Okada, S. Tobishima, J. Yamaki, J. Solid State Ionics 86–88 (1996) 785; Y. Takeda, M. Nishijima, M. Yamahata, K. Takeda, N. Imanishi, O. Yamamoto, Solid State Ionics 130 (2000) 61. [7] R. Alcantara, F.J. Fernandez-Madrigal, P. Lavela, J.L. Tirado, J.-C. Jumas, J. Olivier-Fourcade, J. Mater. Chem. 9 (1999) 2517. [8] M. Lefebvre-Devos, M. Lassalle, X. Wallart, J. Olivier-Fourcade, L. Monconduit, J.-C. Jumas, Phys. Rev. B 63 (2001) 125110; L. Monconduit, M. Tillard-Charbonnel, C. Belin, J. Solid State Chem. 156 (2001) 37. [9] D.C.S. Souza, V. Pralong, A.J. Jacobson, L.F. Nazar, Science 296 (2002) (in press). [10] W. Biltz, M. Heimbrecht, Z. Anorg. Allgem. Chem. 241 (1939) 238. [11] Data were collected on a Siemens D500 equipped with a graphite monochromator using Cu–Ka radiation, and a step size of 0.05° with a counting time of 30 s on each step. [12] M.M. Treacy, NEC, Princeton Labs, Personal communication, 2002. [13] S. Imamura, S. Ishida, H. Tarumoto, Y. Saito, T. Ito, J. Chem. Soc. Faraday Trans. 89 (1993) 757. [14] J. Chastain (Ed.), Handbook of X-ray Photoelectron Spectroscopy, Perkin–Elmer Corporation, Eden Prairie, 1992. [15] G. Nazri, Solid State Ionics 34 (1989) 97. [16] D. Marmorstein, T.H. Yu, K.A. Striebel, F.R. McLarnon, J. Hou, E.J. Cairns, J. Power Sources 89 (2000) 219.